Naturally Dried, Double Nitrogen-Doped 3D Graphene Aerogels

Nov 20, 2017 - ADVERTISEMENT .... Nitrogen-doped graphene aerogels; Oil adsorption; Plectranthus amboinicus; Supercapacitors; Superelasticity...
0 downloads 0 Views 10MB Size
Research Article Cite This: ACS Sustainable Chem. Eng. 2018, 6, 1172−1181

pubs.acs.org/journal/ascecg

Naturally Dried, Double Nitrogen-Doped 3D Graphene Aerogels Modified by Plant Extracts for Multifunctional Applications Qi Meng, Hengcheng Wan, Wenkun Zhu, Tao Duan,* and Weitang Yao* Sichuan Co-Innovation Center for Energetic Materials, Sichuan Civil-Military Integration Institute, Southwest University of Science and Technology, No. 59 Qinglong Avenue, Fucheng District, Mianyang, Sichuan 621010, P. R. China S Supporting Information *

ABSTRACT: Two-dimensional graphene has become one of the most intensively explored carbon allotropes in materials science owing to its attractive features, such as its outstanding physicochemical properties. To advance its practical application, the fabrication of self-assembled 2D graphene sheets into 3D nitrogen-doped graphene aerogels (NGA) with novel functions is becoming essential. Herein, the first attempts to modify graphene with Plectranthus amboinicus (PA), to introduce double nitrogen doping, and to prepare NGA by a natural drying (ND) method are reported. Natural drying was achieved by increasing the pore structure of NDPA-NGA using PA to simultaneously cross-link the graphene sheets and increase the stiffness of the sample. Meanwhile, we used ammonia and urea as double nitrogen sources to achieve an N-doping amount as high as 12.06 atom %. In addition, NDPA-NGA exhibited superelastic properties (with 95% maximal strain and almost no loss after 60 replicates), a high oil absorption capacity, and an excellent electrochemical performance (371 and 204 F g−1 at current densities of 1 and 50 A g−1, respectively, corresponding to a retention of 54.987%). In addition, NDPA-NGA also demonstrated predominant refractoriness, low density, hydrophobicity, and physicochemical stability, which make this material an excellent candidate for use in energy storage, catalysis, and other applications. KEYWORDS: Nitrogen-doped graphene aerogels, Plectranthus amboinicus, Superelasticity, Supercapacitors, Oil adsorption



INTRODUCTION

It is common knowledge that most methods used to prepare 3D NGA employ freeze-drying or vacuum drying,29−31 which requires expensive equipment, thus raising the cost and limiting production. Although previous work has explored methods for naturally drying graphene aerogels (GA),32 the use of chemical reductants and N doping in graphene aerogels (GA) has not been addressed in discussions on the current state of the NGA fabrication process. Obviously, the cost-effective and efficient preparation of NGA is critical for industrial production. In this work, we solved the problem of low NGA doping levels by using ammonia and urea as double nitrogen sources. The nitrogen content was as high as 12.06 atom %, which is close to the theoretical limit of 19.1 atom %.33 Thus, NDPANGA lays the foundation for the preparation of various functional materials.34−37 As previously reported,38 higher natural plants use borate to enhance their intercellular structure by hardening the cell walls to increase their support strength, and the effect of borate on the preparation of bioinspired graphene membranes and graphene aerogels has been studied.39,40 Unfortunately, plant extracts have not been directly used in a more environmentally friendly and NGA-

Since graphene was first successfully prepared by the mechanical stripping method, it has attracted global attention because of its perfect crystal structure and many excellent physicochemical properties.1,2 The unique electrical,3 thermal,4 and mechanical properties5 of graphene have wide application prospects in the fields of electronic devices,6,7 composites,8,9 sensors,10 catalysts,11,12 drug delivery,13,14 and energy storage.15,16 However, the conductivity of graphene cannot be completely controlled because it does not have a bandgap.17 The doping of graphene is carried out by replacing the C atoms with a heteroatom, including N, B, P, and S, to create a dot defect that opens the graphene bandgap and modulates the conductivity type.18−21 Therefore, the elemental doping of graphene, especially with N, has been extensively reported in recent years.22−24 In most reports, however, low N doping has limited the widespread use of nitrogen-doped NGA, and chemicals have been used as modifiers, such as ethylenediamine,25 hexamethylenetetramine,26 layered double hydroxide,27 and L-ascorbic acid.28 These chemicals not only raise the cost of preparation but also cause serious environmental pollution. Therefore, the discovery of new NGA modifiers and methods for preparing NGA with a high N-doping content is a major challenge. © 2017 American Chemical Society

Received: September 26, 2017 Revised: November 6, 2017 Published: November 20, 2017 1172

DOI: 10.1021/acssuschemeng.7b03460 ACS Sustainable Chem. Eng. 2018, 6, 1172−1181

Research Article

ACS Sustainable Chemistry & Engineering

Figure 1. (a) Schematic representation and photo images of the synthesis procedure for NDPA-NGA. (b) Schematic illustration of simultaneous implementation the NDPA-NGA nitrogen-doped, borate cross-linking of graphene sheets and removal of oxygen-containing functional groups (OCFG) by PAPE. cleaning of Plectranthus amboinicus was performed with pure water and anhydrous ethanol. First, 100 g of Plectranthus amboinicus and 300 mL of ultrapure water were added to a juicer. Then, the liquid was removed and centrifuged at 9000 rpm until no particles were present. The slurry was then sonicated for half an hour, and finally, the Plectranthus amboinicus plant extracts were stored in a refrigerator for future experiments (the preparation process is shown in Figure S1, Supporting Information). Graphene oxide was synthesized using a modified Hummers’ method through the oxidation of graphite flakes, as reported elsewhere.41 A GO solution (6 mL, 10 mg/mL) was stirred with 4 mL of concentrated aqueous ammonia (30%) in a beaker. Next, 3 mL of the Plectranthus amboinicus plant extract and 60 mg of urea were added to the solution, followed by ultrasonication to ensure a uniform dispersion. After that, the mixture was transferred to a reaction vessel containing a Teflon liner and maintained at 180 °C for 12 h. The hydrogel was removed and placed in an aqueous solution of ethanol (1%) for 5 h and then frozen in a refrigerator for 10 h. The sample was removed and naturally dried to form NDPA-NGA. Characterization. Scanning electron microscopy (SEM), which gave information about the size and morphology of NDPA-NGA, was performed with a field emission scanning electron microanalyzer (Zeiss Supra 40). X-ray photoelectron spectra (XPS) were recorded on an ESCALab MKΠ X-ray photoelectron spectrometer, using Mg Kα radiation as the exciting source. FTIR spectra were measured on a

suitable preparation method. Inspired by this, we prepared NDPA-NGA from a natural drying (ND) process by using pure natural Plectranthus amboinicus (PA) plant extract (PAPE) as a simultaneous cross-linker and reducing agent. This method can prevent deformation of the NDPA-NGA during the natural drying process and can reduce the oxygen content in NGA. This is a highly competitive mode of production because the PAPE will not pollute the environment, and natural drying is inexpensive, facile, and environmentally friendly. This is the first report of modification using PAPE, the introduction of double nitrogen doping, and the preparation of NGA using natural drying. We firmly believe that this work not only provides a new method for producing nitrogen-doped graphene on a large scale but also presents a new approach for the industrial production of functional N-doped graphene materials.



EXPERIMENTAL SECTION

Synthesis of NDPA-NGA Composites. All chemicals were analytical grade and commercially available from Shanghai Chemical Reagent Co. Ltd. and used as received without further purification. Plectranthus amboinicus is a tender, fleshy perennial plant in the family Lamiaceae and has an oregano-like flavor and odor. Repeated 1173

DOI: 10.1021/acssuschemeng.7b03460 ACS Sustainable Chem. Eng. 2018, 6, 1172−1181

Research Article

ACS Sustainable Chemistry & Engineering Table 1. Theoretical and Actual Volume Retention Ratio of Different Samplesa Sample numbering Drying methods γ (N m−1) 0 1 2 3 4 5

FD, PAPE-3 FD, PAPE-0 ND, PAPE-1 ND, PAPE-3 ND, PAPE-5 ND, PAPE-0

0.073 0.073 0.073 0.073 0.073 0.073

θ (deg)

∼132.93 ∼137.26 ∼129.11

r (μm)

K (KPa)

∼140 ∼140 ∼140

∼162 ∼181 ∼124

Theoretical volume retention ratio (%)

95.622 95.794 94.702

Actual volume retention ratio (%) 98.264 97.879 94.532 95.643 92.141 75.274

a

FD: freeze-drying; ND: Natural drying. PAPE-0, 1, 3, 5 are the amounts that use Plectranthus amboinicus (PA) plant extract (PAPE) as 0, 1, 3, 5 mL, respectively. The frame stiffness (K) is approximated by the elasticity modulus. Theoretical and actual volume retention rates are calculated by eqs 5 and 2, respectively.

Figure 2. (a) SEM image of cross-section view of the 3D-layered porous skeleton structure of NDPA-NGA. (b) SEM image of the core region in the NDPA-NGA. (c) SEM image of local magnification in core region. (d) SEM image of the cross-linking between graphene sheets. (e) SEM images of the cracked cell walls in NDPA-NGA. (f) SEM image of the thicknesses of the cell walls. Bruker Vector-22 FT-IR spectrometer from 4000 to 400 cm−1 at room temperature. The mechanical properties were characterized by a TA Instruments dynamic mechanical analyzer (DMA, RSA3) with compression at ambient conditions. Compressive stress−strain response was measured at a strain rate of 0.05 mm s−1. The specific surface area and the porous features were investigated by nitrogen isothermal adsorption/desorption measurements on a QuadraSorb SI system (Quantachrome Instruments). All the samples were degassed by heating at 150 °C under vacuum for 10 h prior to the measurements. Calculation and Measurements. Calculation of Density and Volume Retention Rate. Sample density (ρ) can be measured by both methods (a) and (b),42 where the average of the data is used.

where V1 is the volume of the sample after 1 h of the hydrogel being placed in the refrigerator; V2 is the volume of the sample after natural drying. Electrochemical measurements. Electrodes were prepared by mixing 80 wt % active materials, 15 wt % carbon black, and 5 wt % polyvinylidene-fluoride in N-methyl pyrrolidone to form the slurry. One square centimeter of this slurry was painted on titanium foil current collectors, and the electrode was dried at 110 °C in a vacuum oven overnight. The mass of active materials in the electrodes was 1−2 mg. Two-electrode (2E) cells were constructed from two carbon electrodes of similar weights, separated by a glassy fiber paper. Threeelectrode (3E) cells were assembled by using one carbon electrode as a working electrode, another carbon electrode as a counter electrode, and inserting an Ag/AgCl reference electrode. Electrochemical performances were collected in 1 M KOH aqueous electrolyte using a Solartron 1480 Multistat station. The test window voltage is −1 to 0 V. The specific capacitance (Cm) is calculated using eq 3:

(a) Calculated by eq 1: ρ = m/(π (d /2)2 h)

(1)

where m is the sample mass; d is the bottom diameter; h is the sample height.

Cm = (I Δt )/(mΔV )

(b) Use the Archimedes principle to directly read the density of the sample through a densitometer. The density of the samples obtained from these two methods is basically the same. Volume retention rate (ζ) is calculated by eq 2:

ζ = V2/V1

(3)

where I is the discharge current; t is the discharge time; V is the voltage change; m is the electrode active material mass. Calculation of Oil Absorption. By measuring the mass m1 and m2 of NDPA-NGA before and after adsorption (the operation should be completed within seconds to avoid volatilization), the oil absorption (Q) was calculated by eq 4:

(2) 1174

DOI: 10.1021/acssuschemeng.7b03460 ACS Sustainable Chem. Eng. 2018, 6, 1172−1181

Research Article

ACS Sustainable Chemistry & Engineering

Figure 3. (a) N2 adsorption/desorption isotherms curves (the amount of PAPE used in sample numbering 2, 3, 4 is 1, 3, 5 mL, respectively). (b) Density and volume retention ratio of different samples (Sample No. 0 is FD with PAPE (3 mL); No.1 is FD without PAPE; No.2 add PAPE 1 mL for ND; No.3 add PAPE 3 mL for ND; No.4 add PAPE 5 mL for ND; No.5 is ND without PAPE). Inset is a contrast photo of No. 3 before and after ND. The density and volume retention are calculated by eqs 1 and 2, respectively.

Figure 4. (a) IR spectra of GO and samples with different PAPE dosages (0, 1, 3, 5 mL). (b) XRD patterns of GO and samples with different PAPE dosages (0, 1, 3, 5, 10 mL). (c) Raman spectra of samples with different urea dosages (0, 30, 45, 60, 75, and 90 mg). Q = m2/m1

and the size of the reactor, and samples of different sizes are shown in Figure S3, Supporting Information). We know that when drying NGA naturally, the evaporation of water produces a capillary force that will cause the 3D structure of the GA to collapse; thus, volume shrinkage and structural cracking are the biggest challenges of natural drying. According to the Laplace formula,44 the capillary force (Δp) is defined as a function of the surface tension (γ), the contact angle (θ), and the hole radius (r), as shown in eq 5:

(4)



RESULTS AND DISCUSSION Figure 1a shows a schematic illustration of the preparation process for NDPA-NGA. First, a GO solution is mixed with ammonia and urea as double nitrogen sources. Then, the modifier PAPE is added, and the mixture is subjected to sonication. During the hydrothermal process, hydrogels are formed by self-assembly as a result of cross-linking between GO and the borate in PAPE (the self-supporting property of the cell wall of tall plants is shown in Figure S2, Supporting Information). In addition, PAPE contains multiple polysaccharide alcohols that replace the oxygen-containing functional groups (OCFG) on the GO sheet,43 which results in a significant reduction of GO; urea and ammonia both act as nitrogen sources during the hydrothermal synthesis process used to dope GA with N atoms. Thus, as shown in Figure 1b, the present method can simultaneously achieve GO reduction, self-assembly and N-doping in one pot. After hydrothermal treatment, the hydrogel is dialyzed to wash away the liquid remaining in the hydrogel while also allowing water molecules to permeate the interior of the hydrogel. The sample is then frozen in a refrigerator to form a macroporous structure in the hydrogel according to the ice template method, after which the sample is removed to form NDPA-NGA at room temperature (the size of the GA prepared using this method is determined by the amount of precursor

Δp = (2γ cos θ )/r

(5)

We forced the graphene layer to extend in the direction of the ice crystal by simple freezing, thereby increasing the pore radius of the GA, and the borate contained in PAPE increased the frame stiffness (K) of the GA so that Δp was much smaller than K. Such a dual effect prevented the volume of the GA from shrinking during the natural drying process (Table 1). As shown in Figure 2 (the TEM image is shown in Figure S4, Supporting Information), NDPA-NGA has a uniformly distributed 3D-layered porous skeleton, which greatly reduces the capillary force. The nitrogen adsorption−desorption isotherm was used to study the specific surface area and pore size. As shown in Figure 3a, the N2 adsorption/desorption isotherms curve of sample numbering 1, 2, 3 can be derived from the different lines corresponding to the relative pressures (P/P0) of 0 to 1.0 (samples numbering 0, 4, 5 are shown in Figure S5, Supporting Information). The typical type-IV 1175

DOI: 10.1021/acssuschemeng.7b03460 ACS Sustainable Chem. Eng. 2018, 6, 1172−1181

Research Article

ACS Sustainable Chemistry & Engineering

Figure 5. (a) XPS overview spectrum of NDPA-NGA and corresponding deconvoluted peaks of (b) N 1s, (c) C 1s, and (d) O 1s spectra. NDPANGA experimental parameter: GO: 6 mL (10 mg/mL); PAPE: 3 mL; ammonia: 4 mL (30%); urea: 60 mg.

the amount of oxygen-containing groups on the surface decreased slightly, suggesting that the hydrothermal treatment without a reductant partially reduced the GO. A much larger decrease in the amount of oxygen-containing groups was recorded in the IR results of the NDPA-NGA samples with different PAPE dosages, indicating that the compounds present in PAPE acted as GO reducing agents. To confirm the effect of PAPE on the GA, the XRD patterns for the different PAPE dosages are shown in Figure 4b. GO exhibits a sharp diffraction peak at 2θ = 10.5°, corresponding to its (002) crystal face. After the hydrothermal reaction, this diffraction peak was no longer observed due to the removal of the oxygen functional groups. The sample showed a broad diffraction peak at 2θ = 25.8° corresponding to the (002) crystal face of the graphite, indicating the presence of stacked graphene nanosheets and the recovery of the graphite crystal structure. The (111) crystal plane of graphite was observed at 2θ = 43°.47 It was found that even when 10 mL of PAPE was added, no miscellaneous peaks appeared in the XRD pattern, indicating the practical use of PAPE for the hydrothermal reduction of GO without introducing impurities. Raman spectroscopy is a commonly used method for evaluating the doping effect of graphene. As shown in Figure 4c, the Raman spectra of GO clearly show two main peaks at 1314 and 1604 cm −1, which correspond to the D and G bands of graphene, respectively. The D band is attributed to vibrations of the sp3 carbon atoms of disordered graphite, and the G band is generated by in-plane vibrations of the sp2 carbon atoms in the 2D hexagonal lattice of graphene. In general, the D/G intensity ratio (ID/IG) reflects the defect density of the carbon material. When 60 mg of urea was used, the maximum ID/IG value of 1.53 was achieved. This result can be attributed to the presence of distortions and vacancies in the graphene lattice caused by N atom doping and to the decrease in the average

features indicate the presence of porous structures. The corresponding pore size was calculated using the BJH method to be 140 μm. In addition, the BET surface area of the samples was 215.11 m2 g−1. The pore size greatly facilitated the drying process of the samples under natural conditions. As shown in Figure 3b, when the sample was freeze-dried, it retained a volume of approximately 97.879%. When 3 mL of PAPE was used in conjunction with the natural drying process, the volume retention was as high as 95.643%, whereas only 75.274% of the volume was retained without PAPE. These results indicate that the addition of PAPE could significantly combat the capillary force (the presence of borate in NDPA-NGA is shown by mapping, Figure S6, Supporting Information). However, upon increasing the amount of PAPE (5 mL), the volume retention decreased, which we attributed to the following reasons: First, there is an optimal amount of borate for the cross-linking of GO. Second, PAPE contains columns of polysaccharides and alcohols that acted as strong reducing agents on GO. The use of a large amount of PAPE hindered the cross-linking of GO by borate and made it difficult to form a perforated structure. In addition, the dilution of the GO precursor by the large amount of PAPE decreased the volume retention rate. The illustration shows the change in the NDPA-NGA volume before and after natural drying with a suitable amount of PAPE (3 mL). In addition, the density of the different samples also changed. PAPE could reduce GO, as mentioned above, which was confirmed using IR spectroscopy. Figure 4a presents the IR spectra of GO, hydrothermally treated GO (without PAPE), and NDPA-NGA with different PAPE dosages. The spectrum of GO contains peaks at 1638, 1389, 1091, 1042, and 878 cm−1, which belong to the CC and NO stretching vibrations of the C−NO groups, the C−N and N−N stretching vibrations of the N−NO groups, and −C−O-C- symmetric stretching, respectively.45,46 After hydrothermal treatment without PAPE, 1176

DOI: 10.1021/acssuschemeng.7b03460 ACS Sustainable Chem. Eng. 2018, 6, 1172−1181

Research Article

ACS Sustainable Chemistry & Engineering

Figure 6. (a) Stress−strain curves at different PAPE dosages (0, 1, 3, 5 mL). (b) Under the 95% strain, the stress−strain curve of different PAPE dosages (0, 1, 3, 5 mL). Inset: Experimental snapshots of one compression cycle. (c) Stress−strain curves of NDPA-NGA at 55% strain for the first five cycles. Inset: Corresponding elastic modulus for the first five cycles. (d) Stress−strain curves of NDPA-NGA at 55%, 85%, and 95% strain for the first 30 cycles. Inset: Variation of height as a function of cycle numbers.

size of the sp2 domain caused by the PAPE-induced reduction of GO. To further characterize the chemical composition of NDPANGA, we performed X-ray photoelectron spectroscopy (XPS) analyses, which showed the characteristic peaks in the C 1s, O 1s, and N 1s spectra (Figure 5a). The highest N doping (maximum of 12.06%) was achieved when 4 mL of the aqueous ammonia solution and a GO:urea mass ratio of 1:1 were used. Table S1 (Supporting Information) details the percentages of C, N, and O obtained from the XPS analysis of each sample. Figure S7 (Supporting Information) contains the XPS images for the different dosages. The NDPA-NGA doping levels were obtained by deconvoluting the N 1s spectrum into three different compositions, which means that the N atoms had three different valences in the bonded graphene,48,49 namely, pyridinic-N (398.7 ± 0.2 eV), pyrrolic-N (399.3 ± 0.2 eV), and graphitic-N (401.4 ± 0.3 eV), as shown in Figure 5b. PyridinicN contributes its p electrons to the graphene layer due to electronic defects, which can improve the initial potential and electrochemical properties of the other N substances in the carbon material, and this property played a decisive role in the NGA sample with a high N content of 30.479%. Pyrrolic-N is an N atom that contributes two p electrons to the graphene π system that is formed from the contribution of pyridinic-N; graphitic-N is derived from the N atom that replaces a C atom in the graphene honeycomb ring, indicating that the N atom is bound to the C−C bonds of graphene. Meanwhile, the highresolution C 1s spectrum could be deconvoluted into five Gaussian components,50−52 as shown in Figure 5c (corresponding to sp2 C, CC, C−C, C−H, 284.4 ± 0.1 eV; sp3 C, C−C, 285.4 ± 0.1 eV; C−N, 286.3 ± 0.1 eV; CN, 287.3 ± 0.1 eV; and CO, OC−O, NC−O, 288.2 ± 0.1 eV). The

presence of a moderate amount of oxygenated groups on the NDPA-NGA skeleton could further decrease the charge transfer resistance by providing more active sites for enhancing power generation. The O 1s peak mainly originated from the physically absorbed oxygen because graphitic carbon is susceptible to oxygen adsorption (the O1 s core level spectra were deconvoluted into four components, corresponding to Ometal; 529.6 ± 0.1 eV; O-physically absorbed, 532.2 ± 0.1 eV; CO, O−CO, 530.8 ± 0.1 eV; and C−OH, C−O−OH, C−O−C, 533.8 ± 0.1 eV),53 as shown in Figure 5d.) The cross-linking effect of the graphene sheets gave NDPANGA excellent elastic properties (Elasticity Video, Supporting Information). The mechanical properties of the samples were tested at a load/unloading rate of 5 mm/min. Figure 6a shows the stress−strain curves for the samples with different PAPE dosages at 55%, 85%, and 95% strain (ε). Although the sample exhibited a large amount of strain (95%), the maximum compressive stress of the sample was quite different (the maximum compressive stress fluctuated from 78 to 180 MPa at 95% strain). When the PAPE dosage was 3 mL, the stress reached 181 KPa, which gave the NDPA-NGA sufficient rigidity to resist the capillary pressure generated during natural drying (Table 1). The loading/unloading curves of all samples at 95% strain are shown in Figure 6b. NDPA-NGA shows different deformation mechanisms for the different compression stages. When ε < 70%, the curve is gentle, and the large pore structure of NDPA-NGA was mainly deformed, while the microstructure remained stable at this stage. When 70% < ε < 95%, the curve shows an obvious increase. During this period, the large pore structure was almost completely compressed, and the microporous, mesoporous, and graphene deformations began to accumulate. The macroscopic performance was 1177

DOI: 10.1021/acssuschemeng.7b03460 ACS Sustainable Chem. Eng. 2018, 6, 1172−1181

Research Article

ACS Sustainable Chemistry & Engineering

Figure 7. (a) CV curves of NDPA-NGA electrode at scan rates from 5 to 50 mV s−1. (b) GCD profiles of NDPA-NGA at current density of 1, 2, 5, and 10 A g−1 and (c) 20, 30, 40, and 50 A g−1. (d) Specific capacitance and rate performance of NDPA-NGA at current density of 1, 2, 5, 10, 20, 30, 40, and 50 A g−1. The specific capacitance and rate is calculated by eq 3.

specific capacitance. Moreover, the specific capacitance of the different nitrogen source ratios at 1 A g−1 was studied using the galvanostatic charge−discharge (GCD) curve, as shown in Figure S8b (Supporting Information (calculated by eq 3). The highest specific capacitance obtained from 60 mg of urea is consistent with the results of the Raman and XPS analyses. We performed a more detailed test on NDPA-NGA-60 to further confirm its electrochemical properties. For the sake of clarity, we still refer to the sample as NDPA-NGA. As shown in Figure 7a, we measured CV curves for NDPANGA at different sweep rates. With the increasing sweep speed, the CVs exhibited no significant changes and maintained a clear rectangle, even at 50 mV s−1, thus showing that the sample had a good rate performance. The CV of NDPA-NGA shows a slightly convex shape in the ranges from −0.8 to −0.6 V and −0.4 to −0.2 V, which was due to the involvement of the electrochemically active functional groups in the redox reaction. The participation of pyrrolic-N was the root cause. Pyridinic-N had a strong effect on the capacitor, while graphitic-N reduced the inherent resistance of the carbon materials and improved their electron transport and capacitance performance under high current loads. Figure 7b and c reflects the GCD curves of NDPA-NGA at different current densities. The specific capacitance was 371 F g−1 at 1 A g−1, which is higher than that in many reports. We speculate that PAPE greatly reduced the amount of oxygen-containing functional groups on graphene so that the N atoms could more easily enter the C atom layer to form carbon−nitrogen bonds. In addition, the specific capacitance of the sample at a current density of 20 A g−1 was still 259 F g−1. Furthermore, the magnification performance of the sample at current densities up to 50 A g−1 was determined. Surprisingly, its capacitance was still 204 F g−1 at this current density, which corresponds to capacity retention of approximately 54.987% and shows the excellent

indicated by a sharp increase in the compressive stress. The illustration shows experimental snapshots from one compression cycle. Figure 6c shows the first five cycles of compression at 55% strain. The stress of the first loading was significantly higher than that of the others and exhibited a maximum elastic modulus of 11.8 KPa, indicating that the stiffness of the sample was sufficient to resist the capillary pressure generated during natural drying. After the first compression, a hysteresis loop is evident, but it gradually decreases as the number of compressions increases. The illustration shows the trend in the modulus of elasticity with the number of cycles. To investigate the trend in the NDPA-NGA recoverable strain with compression and strain, we carried out 10 cycles under 55%, 85%, and 95% strain. As shown in Figure 6d, there is no noticeable difference in the macroscopic morphology of NDPA-NGA after 30 cycles, except the hysteresis loop. The maximum compressive stress underwent the expected contraction and decrease, respectively, and the sample maintained the original height without rupture, indicating that NDPA-NGA could be deformed at any stage (95% or less), demonstrating its wide application. The illustration shows the relative height according to the number of cycles for loading and unloading. The sample at 95% of the maximum strain has a promising future in the production of various structural materials (a comparison between the modifier used, the recoverable maximum strain, and the drying method of NDPA-NGA with those of other graphene aerogel materials is shown in Table S2, Supporting Information). In addition to superelasticity, the N doping of NDPA-NGA enables its application in supercapacitors. The cyclic voltammetry (CV) curve of NDPA-NGA-60 exhibits the largest circumferential area at a scan speed of 10 mV s−1 (Figure S8a, Supporting Information), indicating that it had the largest 1178

DOI: 10.1021/acssuschemeng.7b03460 ACS Sustainable Chem. Eng. 2018, 6, 1172−1181

Research Article

ACS Sustainable Chemistry & Engineering

Figure 8. (a) NDPA-NGA adsorbing gasoline dyed with Sudan III. (b) Water contact measure of NDPA-NGA; the contact angle is 137.26°. (c) Removal of absorbed gasoline by combustion. (d) Attenuation curve of the five kinds of oil after 30 cycles of absorption by NDPA-NGA, where the Q value is calculated by eq 4. Linear fitting equation: edible oil: y = 79.365 − 0.712x; gasoline: y = 35.145 − 0.035x; diesel: y = 31.892 − 0.0764x; pepper oil: y = 29.21 − 0.232x; kerosene: y = 23.379 − 0.036x.

excellent recoverable elastic deformation, which could be completely eliminated by squeezing the adsorbent for its repeated use. NDPA-NGA is also fire resistant, as shown in Figure 8c. Therefore, in this work, we removed the absorbed oil via combustion (Absorption Video, Supporting Information). A total of 30 absorption cycles were conducted with five oils to generate the oil absorption decay curve shown in Figure 8d. In the first cycle, the consumption of edible oil was ∼80 times greater than the initial mass of the material, and after 30 cycles, NDPA-NGA still absorbed ∼58.0 times more oil than its own mass. The mass of absorption of gasoline, diesel, pepper oil, and kerosene was ∼34.1, 29.6, 22.3, and 22.3 times the mass of the absorbent, respectively. The material was able to maintain the original structure after 30 cycles of absorption (Figure S10, Supporting Information), which means that more than 30 cycles of absorption were performed. These results prove that NDPA-NGA is an ideal reusable oil-absorbing material.

rate performance (Figure 7d). This performance is better than that in most other reports (Table S3, Supporting Information). Electrochemical impedance spectroscopy (EIS) is a powerful means of studying the resistance between the electrode and the electrolyte and the internal resistance of the electrode. Figure S9 (Supporting Information) shows the EIS measurements in the frequency range from 0.01 Hz to 1 MHz. The semicircle in the high frequency region represents the charge transfer resistance at the electrode−electrolyte interface in the electrochemical process. In addition, the straight line in the low frequency region represents the ability of the ions to diffuse toward the electrode surface. NDPA-NGA shows a smaller impedance value of approximately 0.73 Ω. This was mainly due to the following reasons: (1) NDPA-NGA had a perfect layerto-pore structure, which facilitated the migration and transport of ions in the electrolyte. (2) The nitrogen doping of graphene increased the electronegativity of the reaction sites, which facilitated the diffusion of the ions on the electrode surface and simultaneously produced more active sites that could participate in the reaction at large current densities. (3) Nitrogen doping prevented restacking between the graphene sheets. Therefore, NDPA-NGA is a promising material for supercapacitors. As shown in Figure 8b, the preparation method did not sacrifice the hydrophobic and lipophilic characteristics of NDPA-NGA; the material had a water contact angle of 137.26° and a porous skeleton arranged in a layered manner that facilitated the rapid absorption of oil, and it provides ideas for the preparation of graphene aerogels with oil absorption properties by natural drying. As shown in Figure 8a, NDPANGA could completely absorb gasoline completely within 5 s and hence has potential applications in various treatments of oil-containing industrial wastewater and marine oil pollution. In addition, NDPA-NGA exhibited a 95% maximum strain and



CONCLUSION

In conclusion, we studied the preparation of NDPA-NGA by natural drying using a pure natural plant extract as a cross-linker and reducing agent. In addition, the use of ammonia and urea as double nitrogen sources was also discussed. The results show that NDPA-NGA exhibited not only excellent volume retention but also excellent elastic and electrochemical properties and oil absorption capacity. This is a solid foundation for the practical application of NDPA-NGA. More noteworthy is that this method is not only the first attempt to prepare NGA but also the first presentation of its unique advantages, such as the ability to introduce high nitrogen-doping content, and it is environmentally friendly, efficient, and eliminates the use of the expensive equipment used in freeze-drying or vacuum drying. This work not only provides a new method for producing 1179

DOI: 10.1021/acssuschemeng.7b03460 ACS Sustainable Chem. Eng. 2018, 6, 1172−1181

Research Article

ACS Sustainable Chemistry & Engineering

(8) Kuilla, T.; Bhadra, S.; Yao, D.; Kim, N. H.; Bose, S.; Lee, J. H. Recent advances in graphene based polymer composites. Prog. Polym. Sci. 2010, 35 (11), 1350−1375. (9) Stankovich, S.; Dikin, D. A.; Dommett, G. H. B.; Kohlhaas, K. M.; Zimney, E. J.; Stach, E. A.; Piner, R. D.; Nguyen, S. T.; Ruoff, R. S. Graphene-based composite materials. Nature 2006, 442 (7100), 282− 286. (10) Shi, X.; Gu, W.; Peng, W.; Li, B.; Chen, N.; Zhao, K.; Xian, Y. Sensitive Pb(2+) probe based on the fluorescence quenching by graphene oxide and enhancement of the leaching of gold nanoparticles. ACS Appl. Mater. Interfaces 2014, 6 (4), 2568−2575. (11) Ba, H.; Podila, S.; Liu, Y.; Mu, X.; Nhut, J. M.; Papaefthimiou, V.; Zafeiratos, S.; Granger, P.; Pham-Huu, C. Nanodiamond decorated few-layer graphene composite as an efficient metal-free dehydrogenation catalyst for styrene production. Catal. Today 2015, 249, 167− 175. (12) Li, Y.; Wang, H.; Xie, L.; Liang, Y.; Hong, G.; Dai, H. MoS2 nanoparticles grown on graphene: an advanced catalyst for the hydrogen evolution reaction. J. Am. Chem. Soc. 2011, 133 (19), 7296− 7299. (13) Weaver, C. L.; Larosa, J. M.; Luo, X.; Cui, X. T. Electrically Controlled Drug Delivery from Graphene Oxide Nanocomposite Films. ACS Nano 2014, 8 (2), 1834−1843. (14) Zhang, L.; Xia, J.; Zhao, Q.; Liu, L.; Zhang, Z. Functional Graphene Oxide as a Nanocarrier for Controlled Loading and Targeted Delivery of Mixed Anticancer Drugs. Small 2010, 6 (4), 537−544. (15) Xiao, L.; Wu, D.; Han, S.; Huang, Y.; Li, S.; He, M.; Zhang, F.; Feng, X. Self-Assembled Fe2O3/Graphene Aerogel with High Lithium Storage Performance. ACS Appl. Mater. Interfaces 2013, 5 (9), 3764− 3769. (16) Zhou, G.; Li, F.; Cheng, H. M. Progress in flexible lithium batteries and future prospects. Energy Environ. Sci. 2014, 7 (4), 1307− 1338. (17) Zhang, Y.; Tan, Y. W.; Stormer, H. L.; Kim, P. Experimental observation of the quantum Hall effect and Berry’s phase in graphene. Nature 2005, 438 (7065), 201−204. (18) Lei, Z.; Lu, L.; Zhao, X. S. The Electrocapacitive Properties of Graphene Oxide Reduced by Urea. Energy Environ. Sci. 2012, 5 (4), 6391−6399. (19) Ding, Y. L.; Kopold, P.; Hahn, K.; van Aken, P. A.; Maier, J.; Yu, Y. Facile Solid-State Growth of 3D Well-Interconnected NitrogenRich Carbon Nanotubedash;Graphene Hybrid Architectures for Lithium Sulfur Batteries. Adv. Funct. Mater. 2016, 26 (7), 1112−1119. (20) Han, J.; Zhang, L. L.; Lee, S.; Oh, J.; Lee, K. S.; Potts, J. R.; Ji, J.; Zhao, X.; Ruoff, R. S.; Park, S. Generation of B-Doped Graphene Nanoplatelets Using a Solution Process and Their Supercapacitor Applications. ACS Nano 2013, 7 (1), 19−26. (21) Su, Y.; Zhang, Y.; Zhuang, X.; Li, S.; Wu, D.; Zhang, F.; Feng, X. Low-temperature synthesis of nitrogen/sulfur co-doped three-dimensional graphene frameworks as efficient metal-free electrocatalyst for oxygen reduction reaction. Carbon 2013, 62 (5), 296−301. (22) Zhang, J.; Li, C.; Peng, Z.; Liu, Y.; Zhang, J.; Liu, Z.; Li, D. 3D free-standing nitrogen-doped reduced graphene oxide aerogel as anode material for sodium ion batteries with enhanced sodium storage. Sci. Rep. 2017, 7 (1), 4886−4892. (23) Wen, Y.; Rufford, T. E.; Hulicovajurcakova, D.; Zhu, X.; Wang, L. Structure Control of Nitrogen-Rich Graphene Nanosheets Using Hydrothermal Treatment and Formaldehyde Polymerization for Supercapacitors. ACS Appl. Mater. Interfaces 2016, 8 (28), 18051− 18059. (24) Deng, D.; Pan, X.; Yu, L.; Cui, Y.; Jiang, Y.; Qi, J.; Li, W. X.; Fu, Q.; Ma, X.; Xue, Q.; Sun, G.; Bao, X. Toward N-Doped Graphene via Solvothermal Synthesis. Chem. Mater. 2011, 23 (5), 1188−1193. (25) Li, J.; Li, J.; Meng, H.; Xie, S.; Zhang, B.; Li, L.; Ma, H.; Zhang, J.; Yu, M. Ultra-light, compressible and fire-resistant graphene aerogel as a highly efficient and recyclable absorbent for organic liquids. J. Mater. Chem. A 2014, 2 (9), 2934−2941.

nitrogen-doped graphene on a large scale but also presents a new approach for the industrial production of N-doped functional graphene materials.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acssuschemeng.7b03460. Absorption Video. (AVI) Elasticity Video. (AVI) Schematic diagram of PAPE preparation, digital image of Plectranthus amboinicus plants, NDPA-NGA test photos and curves, and NDPA-NGA performance comparison table. (PDF)



AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected] (T. Duan). *E-mail: [email protected] (W. Yao). ORCID

Weitang Yao: 0000-0001-9558-0156 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors are grateful for support from the National Natural Science Foundation of China (No. 21671160), Basic Scientific Research Key Project (JCKY2016208B012); Open Foundation of Joint Laboratory for Extreme Conditions Matter Properties, Southwest University of Science and Technology and Research Center of Laser Fusion, CAEP (14tdjk02); Scientific Research Fund of SiChuan Provincial Education Department (11ZB191); and Open Project of State Key Laboratory Cultivation Base for Nonmetal Composites and Functional Materials (12zxnp08). The data were tested by the Southwest University of Science and Technology analytical and testing center.



REFERENCES

(1) Novoselov, K. S.; Geim, A. K.; Morozov, S. V.; Jiang, D.; Zhang, Y.; Dubonos, S. V.; Grigorieva, I. V.; Firsov, A. A. Electric Field Effect in Atomically Thin Carbon Films. Science 2004, 306 (5696), 666−669. (2) Geim, A. K.; Novoselov, K. S. The rise of graphene. Nat. Mater. 2007, 6 (3), 183−191. (3) Orlita, M.; Faugeras, C.; Plochocka, P.; Neugebauer, P.; Martinez, G.; Maude, D. K.; Barra, A. L.; Sprinkle, M.; Berger, C.; de Heer, W. A.; Potemski, M. Approaching the dirac point in high-mobility multilayer epitaxial graphene. Phys. Rev. Lett. 2008, 101 (26), 973− 980. (4) Balandin, A. A.; Ghosh, S.; Bao, W.; Calizo, I.; Teweldebrhan, D.; Miao, F.; Lau, C. N. Superior Thermal Conductivity of Single-Layer Graphene. Nano Lett. 2008, 8 (3), 902−907. (5) Lee, C.; Wei, X.; Kysar, J. W.; Hone, J. Measurement of the elastic properties and intrinsic strength of monolayer graphene. Science 2008, 321 (5887), 385−388. (6) Heo, J.; Oh, J. W.; Ahn, H. I.; Lee, S. B.; Cho, S. E.; Kim, M. R.; Lee, J. K.; Kim, N. Synthesis and characterization of triphenylaminebased organic dyes for dye-sensitized solar cells. Synth. Met. 2010, 160 (19), 2143−2150. (7) Li, Y.; Hu, Y.; Zhao, Y.; Shi, G.; Deng, L.; Hou, Y.; Qu, L. An electrochemical avenue to green-luminescent graphene quantum dots as potential electron-acceptors for photovoltaics. Adv. Mater. 2011, 23 (6), 776−780. 1180

DOI: 10.1021/acssuschemeng.7b03460 ACS Sustainable Chem. Eng. 2018, 6, 1172−1181

Research Article

ACS Sustainable Chemistry & Engineering (26) Moon, I. K.; Yoon, S.; Chun, K. Y.; Oh, J. Highly Elastic and Conductive N-Doped Monolithic Graphene Aerogels for Multifunctional Applications. Adv. Funct. Mater. 2015, 25 (45), 6976−6984. (27) Fang, Q.; Chen, B. Self-assembly of graphene oxide aerogels by layered double hydroxides cross-linking and their application in water purification. J. Mater. Chem. A 2014, 2 (23), 8941−8951. (28) Zhang, X.; Sui, Z.; Xu, B.; Yue, S.; Luo, Y.; Zhan, W.; Liu, B. Mechanically strong and highly conductive graphene aerogel and its use as electrodes for electrochemical power sources. J. Mater. Chem. 2011, 21 (18), 6494−6497. (29) Li, Y.; Chen, J.; Huang, L.; Li, C.; Hong, J. D.; Shi, G. Highly compressible macroporous graphene monoliths via an improved hydrothermal process. Adv. Mater. 2014, 26 (28), 4789−4793. (30) Zhang, Q.; Xu, X.; Lin, D.; Chen, W.; Xiong, G.; Yu, Y.; Fisher, T. S.; Li, H. Hyperbolically Patterned 3D Graphene Metamaterial with Negative Poisson’s Ratio and Superelasticity. Adv. Mater. 2016, 28 (11), 2229−2237. (31) Qiu, L.; Liu, J. Z.; Chang, S. L.; Wu, Y.; Li, D. Biomimetic superelastic graphene-based cellular monoliths. Nat. Commun. 2012, 3 (4), 1241−1247. (32) Xu, X.; Zhang, Q.; Yu, Y.; Chen, W.; Hu, H.; Li, H. Naturally Dried Graphene Aerogels with Superelasticity and Tunable Poisson’s Ratio. Adv. Mater. 2016, 28 (41), 9223−9230. (33) Zhang, S.; Tsuzuki, S.; Ueno, K.; Dokko, K.; Watanabe, M. Upper Limit of Nitrogen Content in Carbon Materials. Angew. Chem., Int. Ed. 2015, 54 (4), 1302−1306. (34) Liao, Y.; Huang, Y.; Shu, D.; Zhong, Y.; Hao, J.; He, C.; Zhong, J.; Song, X. Three-dimensional nitrogen-doped graphene hydrogels prepared via hydrothermal synthesis as high-performance supercapacitor materials. Electrochim. Acta 2016, 194, 136−142. ́ (35) Sliwak, A.; Grzyb, B.; Díez, N.; Gryglewicz, G. Nitrogen-doped reduced graphene oxide as electrode material for high rate supercapacitors. Appl. Surf. Sci. 2017, 399 (31), 265−271. (36) Sui, Z. Y.; Meng, Y. N.; Xiao, P. W.; Zhao, Z. Q.; Wei, Z. X.; Han, B. H. Nitrogen-doped graphene aerogels as efficient supercapacitor electrodes and gas adsorbents. ACS Appl. Mater. Interfaces 2015, 7 (3), 1431−1438. (37) Lee, M. S.; Choi, H. J.; Baek, J. B.; Chang, D. W. Simple solution-based synthesis of pyridinic-rich nitrogen-doped graphene nanoplatelets for supercapacitors. Appl. Energy 2017, 195, 1071−1078. (38) An, Z.; Compton, O. C.; Putz, K. W.; Brinson, L. C.; Nguyen, S. T. Bio-inspired borate cross-linking in ultra-stiff graphene oxide thin films. Adv. Mater. 2011, 23 (33), 3842−3846. (39) Wu, D.; Fu, R.; Zhang, S.; Dresselhaus, M. S.; Dresselhaus, G. Preparation of low-density carbon aerogels by ambient pressure drying. Carbon 2004, 42 (10), 2033−2039. (40) Yang, H.; Zhang, T.; Jiang, M.; Duan, Y.; Zhang, J. Ambient pressure dried graphene aerogels with superelasticity and multifunctionality. J. Mater. Chem. A 2015, 3 (38), 19268−19272. (41) Hummers, W. S., Jr.; Offeman, R. E. Preparation of Graphitic Oxide. J. Am. Chem. Soc. 1958, 80 (6), 1339−1339. (42) Song, X.; Lin, L.; Rong, M.; Wang, Y.; Xie, Z.; Chen, X. Musselinspired, ultralight, multifunctional 3D nitrogen-doped graphene aerogel. Carbon 2014, 80 (1), 174−182. (43) Zheng, Y.; Wang, A.; Cai, W.; Wang, Z.; Peng, F.; Liu, Z.; Fu, L. Hydrothermal preparation of reduced graphene oxide-silver nanocomposite using Plectranthus amboinicus leaf extract and its electrochemical performance. Enzyme Microb. Technol. 2016, 95, 112−117. (44) Zarzycki, J.; Prassas, M.; Phalippou, J. Synthesis of glasses from gels: the problem of monolithic gels. J. Mater. Sci. 1982, 17 (11), 3371−3379. (45) Zhang, J.; Yang, H.; Shen, G.; Cheng, P.; Zhang, J.; Guo, S. Reduction of graphene oxide vial-ascorbic acid. Chem. Commun. 2010, 46 (7), 1112−1114. (46) Li, X.; Wang, Q.; Zhao, Y.; Wu, W.; Chen, J.; Meng, H. Green synthesis and photo-catalytic performances for ZnO-reduced graphene oxide nanocomposites. J. Colloid Interface Sci. 2013, 411 (6), 69−75.

(47) Wang, G.; Yang, J.; Park, J.; Gou, X.; Wang, B.; Liu, H.; Yao, J. Facile Synthesis and Characterization of Graphene Nanosheets. J. Phys. Chem. C 2008, 112 (22), 8192−8195. (48) Wang, H.; Maiyalagan, T.; Wang, X. Review on Recent Progress in Nitrogen-Doped Graphene: Synthesis, Characterization, and Its Potential Applications. ACS Catal. 2012, 2 (5), 781−794. (49) Tian, Z.; Dai, S.; Jiang, D. Stability and Core-Level Signature of Nitrogen Dopants in Carbonaceous Materials. Chem. Mater. 2015, 27 (16), 5775−5781. (50) Roy, S. S.; Mccann, R.; Papakonstantinou, P.; Maguire, P.; Mclaughlin, J. A. The structure of amorphous carbon nitride films using a combined study of NEXAFS, XPS and Raman spectroscopies. Thin Solid Films 2005, 482 (1−2), 145−150. (51) Ishpal; Panwar, O. S.; Kumar, M.; Kumar, S. X-ray photoelectron spectroscopic study of nitrogen incorporated amorphous carbon films embedded with nanoparticles. Appl. Surf. Sci. 2010, 256 (24), 7371−7376. (52) Jiang, L.; Fitzgerald, A. G.; Rose, M. J.; Cheung, R.; Rong, B.; van der Drift, E. X-ray photoelectron spectroscopy studies of the effects of plasma etching on amorphous carbon nitride films. Appl. Surf. Sci. 2002, 193 (1−4), 144−148. (53) Zhu, G. X.; Xi, C.; Liu, Y.; Zhu, J.; Shen, X. P. CN Foam Loaded with Few-layer Graphene Nanosheets for High-performance Supercapacitor Electrode. J. Mater. Chem. A 2015, 3 (14), 7591−7599.

1181

DOI: 10.1021/acssuschemeng.7b03460 ACS Sustainable Chem. Eng. 2018, 6, 1172−1181